MAO-B inhibitors: multiple roles in the therapy of neurodegenerative disorders?

https://doi.org/10.1016/S1353-8020(99)00043-7Get rights and content

Abstract

Monoamine oxidases play a central role in catecholamine catabolism in the central nervous system. The biochemical and pharmacological properties of inhibitors of the monoamine oxidase type B are reviewed. The evidence for biochemical activities distinct from their ability to inhibit MAO-B is discussed, including possible antioxidative and antiapoptotic activities of these agents. The significance of these properties for the pharmacological management of Parkinson's disease and the evidence for a neuroprotective effect of one such agent (selegiline) is also discussed.

Introduction

Monoamine oxidase (MAO; amine: oxygen reductase (deaminating, flavin-containing), EC 1.4.3.4) is an integral protein of the outer mitochondrial membrane, and catalyses the following overall oxidative deamination reaction:RCH2NH2+H2O+O2→RCHO+NH3+H2O2MAO plays a major role in the in vivo inactivation of biogenic and diet-derived amines in both the central nervous system (CNS) and in peripheral neurons and tissues; the most important substrates for the enzyme in the CNS are the catecholamine neurotransmitters (dopamine (DA), adrenaline and noradrenaline (NA)), serotonin (5-hydroxytryptamine; 5HT) and β-phenylethylamine (PE). Two MAO isozymes are distinguished on the basis of their substrate preferences and sensitivity to inhibition by the MAO inhibitor, clorgyline [1]:

  • MAO type A (MAO-A), which is selectively and irreversibly inhibited by low concentrations (nM) of clorgyline. In the human CNS, it is chiefly responsible for the deamination of 5HT and NA; in the intestine, it metabolizes the oxidation of tyramine.

  • MAO-B, which is relatively insensitive to clorgyline. MAO-B inhibition in the human brain principally reduces the catabolism of DA and PE.

The immunologically distinct isozymes are coded for by separate but closely related genes on the X chromosome [2]. The active sites of the two forms exhibit a 93.9% sequence identity [3]; it is the secondary binding sites of the two molecules and possibly their lipid environment which confer their substrate selectivity ([4], [5], [6]; Table 1). Most relevant for our discussion is the fact that DA in the human brain is a substrate for both MAO-A and -B [7].

It is now recognized that there exist significant differences between the rodent and human brains with respect to the regional and cellular distribution of the two MAO forms, as well as the substrate specificity and sensitivity of the two isozymes (Table 2). In contrast to the rat brain, MAO-B is the major form in the human and guinea pig CNS [8], [9], [10], [11], [12]. Moreover, the rat MAO isozymes are reported to have short half-lives, whereas that of MAO-B in primate brain is at least 30 days [13], [14]. These differences are of critical importance, as rodents have been employed for most laboratory studies; such results can only be extrapolated to the human CNS with caution.

Given the substrate specificity of the two MAO forms, their distribution in the human brain is perhaps surprising: the highest MAO-A concentrations are in the catecholaminergic neurons of the locus ceruleus, and of MAO-B in the serotonergic and histaminergic neurons of the raphe and posterior hypothalamus [12], [15], [16], [17]. There are especially high concentrations of both forms in the human basal ganglia [9]. The DAergic neurons of the substantia nigra (SN) in both rodents and primates express MAO-A, but not MAO-B [15]. Although Westlund et al. [16] identified a dense distribution of MAO-B-immunoreactive nerve terminals in the SN, which contrasted with the weak signal for the striatum, nigral MAO-B is located primarily in glial cells [18], [19]. Richards et al. [20] employed quantitative enzyme autoradiography to identify a three-fold higher level of MAO-B than of MAO-A in the SN; however, the levels of the A-form were higher in the pars compacta than in the reticulata, while the opposite distribution was noted for MAO-B.

The topographic location of the MAO types thus does not coincide with that of their presumed natural substrates, and the role of MAO may principally be to protect the local environment from excess levels of foreign monoamines. MAO-B may indirectly regulate extraneuronal transmitter levels, particularly those of DA, by regulating the levels of a release-promoting substance, such as PE [21]. MAO-A, on the other hand, acts to maintain low intraneuronal concentrations of DA, NA and 5HT.

Important for our discussion is the fact that platelet MAO is principally MAO-B (Table 2). As the only means for assessing MAO inhibition in the CNS of a living subject mitigate against their mass application (for example, positron emission tomography (PET) [22], [23]), inhibition of platelet MAO-B activity is the standard parameter employed for estimation of the effect of a MAO-B inhibitor.

CNS MAO-B (but not MAO-A) activity increases with age in both humans and animals [24], [25], [26], [27], [28], perhaps as a result of the glial cell proliferation associated with neuronal loss. In humans, this increase commences at 50–60 years of age, but is not observed in the SN [29]. Increased MAO-B levels in Alzheimer's plaques have also been reported [29], [30]. Increased blood platelet MAO-B activity has been reported in both Alzheimer's (AD) and Parkinson's diseases (PD), although the increased activity in the latter disorder might be associated with the frequently coexistent AD, rather than directly with PD [31]. Fowler et al. [32] reported that MAO-B activity was reduced by 40% in the brains of smokers; tobacco use is associated both with psychiatric disease and with a reduced risk for PD.

Section snippets

MAO-B inhibitors

MAO inhibitors were employed clinically in the 1950s as antidepressants, but fell out of favour because of the so-called “cheese effect”. These first MAO inhibitors not only had the desired effect of elevating CNS catecholamine levels, but also potentiated the sympathomimetic action of indirectly acting amines, including tyramine (an MAO-A substrate), in the periphery; hypertensive crises following the consumption of tyramine-containing foods, such as cheese and red wine, were thus a dangerous

Inhibition of MAO-B

Interest in the MAO-B inhibitors was initially stimulated by the desire to elevate the reduced striatal DA concentrations characteristic of PD. As MAO-B is generally present in excess in the tissues in which it occurs, it is necessary to inhibit at least 80% of the enzyme to achieve a pharmacological effect [86]. A daily dose of 10 mg selegiline achieves total inhibition of platelet MAO-B [64], [87]. In phase I trials, rasagiline achieved 90% inhibition of platelet MAO-B at a dose of 1 mg/day for

Clinical aspects of MAO-B inhibitors

As selegiline is the only MAO-B inhibitor currently licenced for the therapy of PD and the cognitive aspects of AD (although a number of other agents (rasagiline, lazabemide, milacemide) are undergoing clinical trials to determine their suitability for the treatment of these and other neurodegenerative diseases; [129], [226]), the following discussion will necessarily focus on this agent. As discussed above, selegiline, originally designated deprenyl, was developed as an antidepressant before

Conclusions

Interest in the development of more specific MAO-B inhibitors was sparked by the success of selegiline as an adjunct to the l-DOPA therapy of PD. Its effects may include both symptomatic and neuroprotective components, although the latter aspect has thus far not been decisively demonstrated. In preclinical studies, selegiline has exhibited a range of properties which appear to go beyond enzyme inhibition (Table 6); the further investigation of alternative MAO-B inhibitors, and the comparison of

References (285)

  • J. Saura et al.

    Increased monoamine oxidase B activity in plaque-associated astrocytes of Alzheimer brains revealed by quantitative enzyme radioautography

    Neuroscience

    (1994)
  • K. Magyar et al.

    Structure–activity relationships of selective inhibitors of MAO-B

  • D.D. Truong et al.

    Milacemide increases 5-hydroxytryptamine and dopamine levels in rat brain—possible mechanisms of milacemide antimyoclonic property in the p,p′-DDT-induced myoclonus

    Pharmacology and Biochemistry of Behavior

    (1989)
  • E.M. O'Brien et al.

    The interactions of milacemide with monoamine oxidase

    Biochemical Pharmacology

    (1994)
  • E.M. O'Brien et al.

    Species differences in the interactions of the anticonvulsant milacemide and some analogues with monoamine oxidase-B

    Biochemical Pharmacology

    (1995)
  • D.A. Di Monte et al.

    Monoamine oxidase-dependent metabolism in the striatum and substantia nigra of l-DOPA-treated monkeys

    Brain Research

    (1996)
  • P.C. Waldmeier et al.

    Deprenil, loss of selectivity for inhibition of B-type MAO after repeatessd treatment

    Biochemical Pharmacology

    (1978)
  • L.M. Kochersperger et al.

    Assignment of genes for human monoamine oxidase A and B to the X chromosome

    Journal of Neuroscience Research

    (1986)
  • J. Grimsby et al.

    Human monoamine oxidase A and B genes exhibit identical exon–intron organization

    Proceedings of the National Academy of Sciences USA

    (1991)
  • A. Kalir et al.

    Selected acetylenic suicide and reversible inhibitors of monoamine oxidase types A and B

    British Journal of Pharmacology

    (1981)
  • J.C. Shih et al.

    Determination of regions important for monoamine oxidase (MAO) A and B substrate and inhibitor sensitivities

    Journal of Neural Transmission

    (1998)
  • C.J. Fowler et al.

    Deamination of dopamine by monoamine oxidase-A and -B in the rat and in man

  • R.F. Squires

    Multiple forms of monoamine oxidase in intact mitochondria as characterized by selective inhibitors and thermal stability: a comparison of eight mammalian species

    Advances in Biochemical Psychopharmacology

    (1972)
  • L. Oreland et al.

    Monoamine oxidase activity and localization in the brain and the activity in relation to psychiatric disorders

  • A.J. Azzaro et al.

    Guinea pig striatum as a model of human dopamine deamination: the role of monoamine oxidase isoenzyme ratio, localization, and affinity for substrate in synaptic dopamine metabolism

    Journal of Neurochemistry

    (1985)
  • S.B. Ross

    Distribution of the two forms of monoamine oxidase within monoaminergic neurons of the guinea pig brain

    Journal of Neurochemistry

    (1987)
  • C.D. Arnett et al.

    Turnover of brain monoamine oxidase measured in vivo by positron emission tomography using l-[11C]deprenyl

    Journal of Neurochemistry

    (1987)
  • J.S. Fowler et al.

    Slow recovery of human brain MAO-B after l-deprenyl (selegiline) withdrawal

    Synapse

    (1994)
  • K.N. Westlund et al.

    Distinct monoamine oxidase A and B cell populations in primate brain

    Science

    (1985)
  • K.N. Westlund

    The distribution of monoamine oxidases A and B in normal human brain

  • J.G. Richards et al.

    Monoamine oxidases: from brain maps to physiology and transgenics to pathophysiology

    Journal of Neural Transmission

    (1998)
  • M.B.H. Youdim

    Platelet monoamine oxidase B: use and misuse

    Experientia

    (1988)
  • C.J. Fowler

    Enzyme activity: monoamine oxidase

  • J.S. Fowler et al.

    Visualization of monoamine oxidase in human brain

    Advances in Pharmacology

    (1998)
  • M. Strolin-Benedetti et al.

    Monoamine oxidase, brain ageing and degenerative diseases

    Biochemical Pharmacology

    (1989)
  • M.D. Galva et al.

    Effect of aging on lazabemide binding, monoamine oxidase activity and monoamine metabolites in human frontal cortex

    Journal of Neural Transmission

    (1995)
  • W. Danielczyk et al.

    Platelet MAO-B activity and the psychopathology of Parkinson's disease, senile dementia and multiinfarction dementia

    Acta Psychiatrica Scandinavica

    (1988)
  • J.S. Fowler et al.

    Inhibition of monoamine oxidase B in the brains of smokers

    Nature

    (1996)
  • B. Blackwell et al.

    Hypertensive interactions between monoamine inhibitors and foodstuffs

    British Journal of Psychiatry

    (1967)
  • K.F. Tipton et al.

    The kinetics of monoamine oxidase inhibitors in relation to their clinical behaviour

  • J. Gaál et al.

    Medicinal chemistry of present and future MAO-B inhibitors

  • R.R. Ramsay

    Substrate regulation of monoamine oxidases

    Journal of Neural Transmission

    (1998)
  • R.H. Abeles et al.

    Suicide enzyme inactivators

    Accounts of Chemical Research

    (1976)
  • J. Knoll et al.

    Some puzzling pharmacological effects of monoamine oxidase inhibitors

    Advances in Biochemical Psychopharmacology

    (1972)
  • J. Knoll et al.

    Phenylisopropylmethyl-propinylamine (E-250), a new spectrum psychic energizer

    Archives of International Pharmacodynamics

    (1965)
  • M.J. Parnham

    The history of l-deprenyl

  • K. Magyar et al.

    Comparative pharmacological analysis of the optical isomers of phenylisopropylmethylpropinylamine (E-250)

    Acta Physiologica Academy of Sciences (Hungary)

    (1967)
  • C.F. Huebner et al.

    N-methyl-N-2-propynyl-1-indanamine. A potent monoamine oxidase inhibitor

    Journal of Medicinal Chemistry

    (1966)
  • J.P.M. Finberg et al.

    Tyramine antagonistic properties of AGN 1135, an irreversible inhibitor of monoamine oxidase type B

    British Journal of Pharmacology

    (1981)
  • M.B.H. Youdim et al.

    MAO type B inhibitors as adjunct to l-dopa therapy

    Advances in Neurology

    (1986)
  • Cited by (86)

    • A pilot reverse virtual screening study suggests toxic exposures caused long-term epigenetic changes in Gulf War Illness

      2022, Computational and Structural Biotechnology Journal
      Citation Excerpt :

      Our observed hypomethylation of the MAOB gene may be a direct response to this inhibition indicative of increased expression to compensate for pesticide exposure. As many drugs have been developed to inhibit MAOB as a neuroprotective treatment to neurodegenerative disorders, such as Alzheimer’s disease, Parkinson’s disease, and ALS [61], increased expression may lead to the opposite effect and increase the likelihood of these illnesses. In support of this finding recent mouse models of GWI have shown that a combination of PB, DEET, permethrin and the sarin analog DFP cause serotonergic dyshomeostasis, monoamine disbalance, and neuroinflammation in multiple brain regions, as well as acute peripheral serotonin imbalance [62].

    • Privileged scaffolds as MAO inhibitors: Retrospect and prospects

      2018, European Journal of Medicinal Chemistry
    • Benzyloxynitrostyrene analogues – A novel class of selective and highly potent inhibitors of monoamine oxidase B

      2017, European Journal of Medicinal Chemistry
      Citation Excerpt :

      In addition, high or repeated concentrations of irreversible MAO-B inhibitors lead to a loss of MAO-B selectivity [17] and thus contribute to the potential adverse effect of tyramine-induced hypertension (“cheese-reaction”) that is often associated with irreversible MAO-A inhibition. On the other hand, reversible, competitive inhibitors bind non-covalently to the active site of the enzyme [16], thus allowing for competition with the substrate. Two irreversible MAO-B inhibitors, rasagiline (IC50 = 0.070 μM) and selegiline (IC50 = 0.020 μM), are currently in use for the treatment of Parkinson's disease [18,19], while a reversible inhibitor, safinamide (IC50 = 0.080 μM), has entered clinical trials [1,19].

    View all citing articles on Scopus
    View full text